You are on page 1of 13

1.

Be Aware of the Fine Print' In the


Science of Metallurgy of Induction
Hardening: Part 2
May 10, 2005
It is important to look closer at the "fine points" of certain metallurgical phenomena and
principles of heat treatment in the application of induction hardening to ensure optimal results.
We are reminded in many aspects of our lives today to read the "fine print" contained in
manuals, warrantees, instructions, etc., so that we don't miss important information. Important
information also is often overlooked in the science of metallurgy. While there is no fine print,
there are metallurgical "fine points" and subtleties, which require paying close attention. In many
cases, certain metallurgical phenomena and principles of heat treatment are incorrectly assumed
or improperly used in induction hardening. It is important to look closer at the "fine points" of
metallurgical theory to ensure optimal results. To illustrate this, following are additional
examples to those discussed in Part 1 of this article (March 2005 IH).

Fig. 1. Influence of carbon content on Ms and Mf temperatures. Source: top (Ref. 4,5); bottom
(Ref. 6).

Is it always possible to obtain a fully martensitic structure of


noneutectoid plain carbon steels using severe water spray
quench?
In hardening applications, the ability to obtain a certain degree of martensitic structure is often
the measure of how successful the heat treating process was. Martensite is a supersaturated solid
solution of carbon in ferrite with a body-centered tetragonal crystalline (BCT) structure. Upon
rapid cooling, carbon is trapped in the crystal structure. The high hardness developed when steel
is transformed to martensite is due to distortion in the transformation from FCC (austenite) to
BCT martensite crystalline structure [1-3]. Martensite formation is governed by a shear-type
(diffusionless) transformation of austenite; that is, the transformation occurs almost
instantaneously upon reaching a certain temperature.

If the continuous cooling transformation (CCT) diagram of steel is shifted far to the left, the
cooling curve will enter the upper transformation start region regardless of quench severity,
preventing the capability to obtain an entirely martensitic microstructure. In such cases, the final
microstructure of the hardened layer will consist of a combination of martensite and upper
transformation products (e.g., pearlite and banite) regardless of the severity of the quench from
austenitic temperature down to ambient temperature. Although a fully martensitic structure might
not be obtained, the amount of upper transformation products can be appreciably small and
might not noticeably affect component mechanical properties in certain applications.
Martensitic transformation occurs over a temperature range between the Ms (martensite start) to
Mf (finish), which depends on the particular steel chemical composition, and from practical
perspective, cannot be changed by varying quench severity. If cooling upon quenching is
interrupted at a certain temperature within the martensite transformation range, no further
transformation to martensite will occur. Martensite transformation resumes upon further cooling
to a lower temperature.
Figure 1 shows that Ms-Mf temperature range is directly related to the steel's carbon content.
The carbon content and actual amount of martensite formed exclusively determine the maximum
hardness of given steel. In the range of 0.2 to 0.65% carbon, the hardness of the steel is
proportional to the carbon content.

Fig. 2. Influence of carbon content on volume of retained austenite


Mf temperatures for plain carbon steels with high carbon content, cast irons and some alloy
steels are well below room temperature. Thus, even if quench severity is sufficient to miss the
upper transformation region of the CCT diagram, a fully martensitic structure will not be
obtained within the hardened layer. The existence of a noticeable amount of untransformed, or
retained, austenite will be unavoidable (Fig. 2). Cryogenic treatment can be used to transform the
retained austenite into martensite. Note that a full untempered martensitic structure has low
ductility [1].
The inability to obtain a fully martensitic structure is directly related to not having homogeneous
austenite. As mentioned in Part 1 of this article, rapid induction heating can result in
inhomogeneous austenite with a ferritic/pearlitic network (dependent on chemical composition,

prior microstructure and process specifics), which remains in the as-quenched structure,
preventing a fully martensitic structure.
Nonuniform distribution of carbon is another characteristic of inhomogeneous austenite. After
fast heating, a ferrite/pearlite network might not exist, but the austenite phase could consist of
localized regions of enriched carbon and reduced carbon. Since both Ms and Mf temperatures
depend on carbon content, austenite areas of high and low carbon concentration will have
different critical cooling rates (CCT curves) resulting in different transformations to martensite.
This may lead to the appearance of alternative products within the martensitic structure.
Carbon content not only influences achievable surface hardness and case depth, but also the
transition zone. For example, eutectoid steel always has a shorter transition zone compared with
a hypoeutectoid steel (assuming temperature distribution and quenching conditions are the
same).

Fig. 3. Influence of test specimen diameter on hardness profile and core cooling rate [1]

Can results of Grossmann's hardenability tests and Jominy


end-quench tests be directly applied to selective hardening?
Hardenability is an important property of steel and cast iron, defining the ability of the metal to
be hardened to a certain depth. It is measured as the distance from the surface where certain
hardness can be obtained or a specific percentage of martensite can be formed (e.g., 50 HRC or
50% martensite). When discussing hardenability, it is important to recognize factors that
influence hardness distribution in through hardening compared with surface hardening.

Fig. 4. Compact system for hardening cast iron camshafts. Courtesy of Inductoheat Inc.
Through hardening (hardening a workpiece through its entire cross section) typically requires
heating the part uniformly to the austenitizing temperature range and quenching to ambient
temperature. During quenching, the cooling rate at the surface always is more intense compared
with the rate at internal locations; particularly compared with the rate at its core. For a thin part,
the cooling intensity at the core might be severe enough to miss the upper transformation region
of the continuous cooling diagram and to form a sufficient amount of martensite in core,
resulting in a relatively uniform through-hardened pattern. Also, because the cooling rate at the
surface is always more intense than that at the core, more martensite forms at the surface and
subsurface than in the core, resulting in a higher hardness at the surface (assuming that surface
has not been overheated and severe oxidation and/or decarburization did not occur).

At larger diameters or thicknesses, the depth of the hardened layer (hardened depth) also
increases. In addition, because the core is at a greater distance from the quenched surface,
thermal conductivity provides less intense cooling of the core during surface quenching. At a
certain point, the CCT curve representing core cooling conditions during surface quenching
shifts farther to the right (Fig. 3), eventually passing through an area where transformation starts
at a temperature higher than the Ms temperature. Thus, depending on the cooling rate, a certain
amount of upper-transformation products (i.e., bainite, pearlite, ferrite) form within the core
leading to a softer core compared with the surface.
Surface hardening (case hardening) involves a relatively short heat up time and a pronounced
skin effect, and core temperature does not rise significantly during the heating stage [1]. Upon
quenching, the cold core provides an additional cooling effect on the case hardening area
compared with through-heated parts. A more intensive quench increases hardness and steel
hardenability, and dramatically increases the formation and distribution of residual stresses [1].

Fig. 5. Representative gray cast iron micro-structure containing graphite flakes [9]; Fig. 6.
Unitized machine with two independently operated heat stations and high-speed, servo-driven
scanning assemblies for induction hardening the ID of gray iron cylinder liners. Production rate
is 50 liners/h with a 0.8 mm (0.03 in.) case having 47-49 HRC as-quenched hardness. Courtesy
of Inductoheat Banyard Ltd., Dorset, UK.
Hardenability tests

Techniques used to determine steel and cast-iron hardenability include Grossmann's


hardenability test and the Jominy end-quench test-the most widely used tests [1-3]. Grossmann's
test involves the concept of critical diameter. Cylindrical bars having different diameters are
cooled from a specific austenitizing temperature down to ambient temperature using a given
quench medium. For a particular quenchant, a bar having 50% martensite at its core would
correspond to the critical diameter Dcr. The critical diameter is influenced by chemical
composition, grain size and homogeneous austenite, which can be relatively easily defined.
However, quenching condition often is the least defined factor, particularly in the case of
induction hardening. Cooling severity during spray quenching depends on a combination of
factors including quenchant type and purity; concentration; quenching temperature, pressure and
flow rate; quench block design; number and distribution of quench holes, orifice size and
density; spray impingement and part rotation.
The Jominy end-quench was developed to overcome the complexity of the Grossmann's
hardenability test. In the test, a 25 mm OD by 100 mm long (~1 in. by 4 in.) cylindrical specimen
is uniformly heated to achieve homogeneous austenite and spray quenched on one end,
producing a longitudinal hardness distribution as a function of the distance from the quenched
end. Both test methods require some degree of caution for induction hardening (particularly
induction surface hardening) due to the assumptions in measurements that have been done.
The standard Jominy end-quench test is suitable for moderate cooling rates. Therefore, it can
provide misleading results at cooling rates 150C/sec (270F/sec) or higher experienced with a
majority of induction hardening applications. Quench severity exceeds 1000C/sec (1830F/sec)
in some induction hardening applications.
In standard hardenability tests, a specimen is heated to the austenitic temperature and held long
enough to ensure forming homogeneous austenite. By comparison, induction hardening involves
intense heating with a relatively short or no holding time, which can produce inhomogeneous
austenite with corresponding differences in the hardenability curves. Intense heating during
induction hardening shifts the A1 and A3 critical curves toward higher temperatures [1,7].
Quenching from temperatures that are often 100 to 180C (180 to 325F) higher than the
temperature used during hardenability tests can result in appreciable errors.
Conventional hardenability curves also are modified by the cold core serving as a heat sink and
by higher surface and subsurface quenching severity. The cold core can have a self-quenching
(mass quenching) effect, which allows the elimination of liquid spray quenching in some
hardening applications with a small case depth.
Hardenability tests are primarily oriented toward cylindrical shape specimens versus other
shapes, which make it difficult to apply test results to parts having complex shape. Data obtained
from a standard Jominy or Grossman test should be used for reference purposes only.

Fig. 7. Induction hardened ductile cast iron crankshaft with corresponding microstructures of the
hardened case, transition zone and green core. Required case depth: 1.8 mm (0.07 in.). Courtesy
of Inductoheat Inc.

Are the procedures for choosing process parameters for


induction hardening of steels and cast irons identical?
One of the most common applications of induction heat treatment is hardening cast iron parts
such as camshafts, crankshafts, liners, gears, rollers, etc. Figure 4 shows an induction system
used to heat treat cast iron camshafts. Induction surface hardening of cast irons has many
similarities to induction hardening of carbon steels, but there are some significant differences [1].
Cast irons comprise a family of materials represented on the right side of the Fe-FeC3 phase
diagram having a high carbon content (2+%) and a wide range of properties [1,7]. Gray, ductile
(nodular) and malleable and compacted-graphite iron (to a lesser extent) are induction hardened.
Gray iron contains carbon as graphite in flake form (Fig. 5), which combined with the high
carbon content makes gray iron castings brittle and hard, with low tensile strength and a poor
ability to withstand appreciable thermal shock and shock load. Due to their relatively high silicon
contents, commercial cast irons should be considered as at least ternary Fe-C-Si alloys. As a
result, all critical temperatures of cast irons differ from those shown in the right side of the FeFe3C diagram [1].
The ability of gray irons to be hardened depends on the type of matrix (i.e., ferritic, ferriticpearlitic or pearlitic) and the amount, size, shape, and distribution of graphite flakes. A pearlitic
matrix provides a better response to induction hardening of gray irons. Being brittle, gray iron
may present certain challenges to induction hardening due to a tendency toward cracking from
fast heating and intense cooling [1,9,10]. Preheating and soft quenching are often used. However,

some gray iron parts have been successfully hardened using short heat time (less than 2 seconds)
and quenching using plant water (Fig. 6).
In contrast to gray irons, ductile irons have carbon particles in the form of graphite nodules,
which serve as crack arresters. This gives ductile irons some advantages over gray irons
including ductility, relatively high tensile and bending strength and moderate elongation. An
induction hardened ductile (nodular) cast iron crankshaft with case hardness pattern and
microstructure, transition zone and green core is shown in Fig. 7.

Ductile (nodular) irons offer a wide range of properties. Five subgroups of ductile iron are
ferritic, pearlitic-ferritic, pearlitic, martensitic and austempered ductile irons. Induction
hardening is usually applied to martensitic, pearlitic and, to a lesser extent, pearlitic-ferritic
ductile irons. Martensitic ductile iron requires the lowest hardening temperatures, shortest heatup time and provides well-defined, crisp hardness patterns with a relatively shallow transition
zone [1].
Being inherently strong, ductile irons can handle much greater stresses than gray irons upon
heating and quenching without cracking. However, the presence of graphite nodules as crackarresters does not guarantee ductile iron castings will not crack during intensive heating or/and
severe quenching.
A temperature range of 860 to 960C (1580 to 1760F) is typical for induction hardening of iron
castings. Besides carbon and silicon, all commercial cast irons also have other intentionally
added alloying elements and residual impurities that could affect critical temperatures.
A key metallurgical difference between steels and cast irons that causes many problems in
selective hardening is matrix carbon content versus total carbon content (CE) [8]. In steels,
carbon content is fixed by chemical composition and cannot exceed this value in a fully
austenitic condition. In contrast, cast irons have a carbon "reserve" in the primary graphite
particles, which can cause localized increasing amounts of carbon to dissolve into the austenite
matrix at higher austenitizing temperatures. This high (and variable depending on temperature)
matrix carbon content is the most critical metallurgical factor in the selective hardening of cast
irons, and lies at the root of problems such as excessive retained austenite, coarse martensite and
unusual hardness patterns. Alloying elements can affect hardenability and retained austenite (by
decreasing Ms temperature) in steels, but matrix carbon content in cast irons have an
overwhelming effect.
Ferritic ductile iron typically is not a good candidate for induction hardening due to the inability
to transform low carbon containing ferrite into martensite. However, intense induction heating

and quenching can be used to improve the fatigue strength of ferritic ductile cast irons [12] from
the ability to create localized martensitic areas and compressive stresses near the boundaries of
graphite nodules due to carbon diffusion from graphite nodules into ferrite matrix.

Fig. 9. Influence of carbon content on electrical resistivity and magnetic permeability of carbon
steels and cast irons at ambient temperature
Sections of varying thickness in complex shaped iron castings heat up at different rates,
promoting thermal gradients and thermal stresses, which can result in distortion and crack
development, particularly in locations having a drastic change of mass. Complications can arise
when transitional thermal stresses combine with residual stresses from previous operations (e.g.,
casting, machining, honing, surface peening, etc.). Complex-shaped steel parts have less
tendency to crack than complex cast iron parts. Stress relieving iron castings prior to induction
hardening is often recommended to reduce the probability of cracking. Formation of stresses
during induction hardening and tempering is discussed in [1]. Reasons for crack development
during hardening cast irons and plain carbon or alloy steels are different [1,9,10]. For example,
age strengthening can occur in gray iron castings but not in steel parts. If age strengthening
occurs, some castings may harden relatively easily, while others may crack, even though heating
and quenching conditions were identical.
A study of the age strengthening phenomenon [11] reported that aging at room temperature for
about 60 days can strengthen gray iron castings by as much as 12%. The tensile strength-tohardness ratio also increases because the hardness does not change with time. In a production
environment, the time between casting and heat treating can be relatively short, and age
strengthening will not occur. Thus, to ensure the reliability and repeatability of a gray iron
hardening operation, it is important to conduct a run-off using relatively "fresh" castings. Using
castings that have been on the shop floor for some time for process development or run-off could
result in hardening age-strengthened parts. Such results could be overly optimistic, and cracking
might suddenly occur during a production run [1].
The response of cast irons to electromagnetic heating is different than that of steels. It is
important to remember that in contrast to alternative heating processes, the intensity of induction

heating is more sensitive to a chemical composition, and is directly related to the electromagnetic
properties of the heated metal.
Electromagnetic properties of materials in a broad sense include magnetic permeability,
electrical resistivity (electrical conductivity), saturation flux density, coercive force, hysteresis
loss, permittivity, magnetic susceptibility and others. While all electromagnetic properties are
important, this discussion is limited to the effect of electrical resistivity and relative magnetic
permeability on the ability of metal to be heated by induction. Electrical resistivity and magnetic
permeability have the most pronounced effect on performance of the induction heating system
(Fig. 8).

Fig. 10. Magnetization B-H curves for cast steel and cast iron [13]
Electrical resistivity, the reciprocal of electrical conductivity, affects nearly all induction heating
system parameters including depth of heating (current penetration depth), heat uniformity, coil
electrical efficiency, coil impedance (load matching capability) and others. Electrical resistivity
varies with temperature, chemical composition, metal microstructure and grain size. It increases
nonlinearly with temperature for steels and cast irons.
Relative magnetic permeability (r) is a nondimensional parameter that indicates the ability of a
material to conduct the magnetic flux better than vacuum or air. It has a marked effect on all
basic induction phenomena including the skin effect, end effect and proximity effect, and also
has a major effect on coil electrical parameters [1]. The magnetic permeability of a particular
metal is a function of both temperature and magnetic field intensity.
Electrical resistivity and magnetic permeability are strongly dependent on the chemical
composition (Figs. 9, 10 and 11). Cast irons have higher electrical resistivity but lower magnetic
properties compared with carbon steels. Thus, a coil using the same power supply and frequency
has a different heating effect on carbon steel, alloy steel or cast iron parts of the same geometry
[1].

Fig. 11. Influence of small amounts of alloying elements on the electrical resistivity of iron [14]
A material's thermal properties also are a function of chemical composition. Since the cycle time
of induction heating is much shorter than those of alternative heating processes, the variation in
thermal conductivity of the heated material has a greater effect on transient and final thermal
conditions of the inductively heated parts.
Thermal conductivity of cast irons is typically lower than that of carbon steels resulting in much
weaker "soaking" action during surface hardening or selective hardening. Therefore, in contrast
to surface hardening of steels, self-quenching is practically never used in hardening cast irons.
For example, in the case hardening of gray iron cyclinder liners in Fig. 6, self-quenching was not
applied even for the required shallow case depth of 0.8 mm (0.03 in.). It is important to
remember that the size, shape, dispersion and amount of graphite flakes affect not only the
mechanical properties, but also the electrical, magnetic and thermal properties of gray cast irons
[1].
When discussing induction surface hardening of steel, the phenonmenon of super hardening is
often mentioned [1,4], wherein the surface hardness of an induction hardened steel could be 2-3
HRC higher than that for through-heated, furnace-hardened steel [1,4]. This phenomenon is
particularly noticeable in induction hardened steels having a 0.35-0.6% carbon content. Super
hardening has never been observed in induction hardening iron castings.

Conclusion
The material discussed in Part 1 and Part 2 of this article is representative of metallurgical "fine
points" and subtleties that sometimes are incorrectly assumed or improperly used in the
application of induction hardening.. However, there many other principles of heat treatment that
must be carefully thought out when applying them to the induction hardening compared with
alternative heat treating methods to prevent unanticipated surprises. IH

References

Rudnev, V., Loveless,D., et al., Handbook of Induction Heating, Marcel Dekker, 2003

Brooks, C., Principles of the Heat Treatment of Plain Carbon and Low Alloy Steels,
ASM Intl., 1996
Krauss, G., Steels: Heat Treatment and Processing Principles, ASM Intl., 1999
Semiatin, S.L. and Stutz, D.E., Induction Heat Treatment of Steel, ASM Intl., 1986
Troiano, A. and Greninger, A., Metal Progress, 1946
Gulyaev, A. Metallurgy, Metallurgia, Moscow, 1977
Rudnev, V., Be aware of the 'fine print' in the science of metallurgy of induction
hardening, Part 1, Ind. Htg., Mar., 2005
Private communication with Norman Carter, May, 2003
Rudnev, V., Induction Hardening Cast Iron, Ht. Trtg. Prog., ASM Intl., Mar., 2003
Rudnev, V., Troubleshooting Cracking in Induction Hardening, Ht. Trtg. Prog., Aug.,
2003
Nicola, W. and Richards, V., Age Strength-ening of Gray Iron, AFS Trans., 2000
Misaka, Y., et. al., Fatigue strength of ferritic ductile cast iron, Japan IOM, 2004
Attwood, S., Electrric and Magnetic Fields, John Wiley, 1941
Bozorth, R., Ferromagnetism, IEEE Press, N.Y., 1993

You might also like